Sie sind auf Seite 1von 31

Annu. Rev. Plant Biol. 2004. 55:37399 doi: 10.1146/annurev.arplant.55.031903.141701 Copyright c 2004 by Annual Reviews.

All rights reserved First published online as a Review in Advance on January 12, 2004

REACTIVE OXYGEN SPECIES: Metabolism,


Oxidative Stress, and Signal Transduction
Klaus Apel1 and Heribert Hirt2
1

Annu. Rev. Plant Biol. 2004.55:373-399. Downloaded from arjournals.annualreviews.org by CAPES on 12/19/08. For personal use only.

Institute of Plant Sciences, Swiss Federal Institute of Technology (ETH) Universit tstr. 2, 8092 Z rich, Switzerland a u 2 Max F. Perutz Laboratories, University of Vienna, Gregor-Mendel-Institute of Molecular Plant Sciences, Austrian Academy of Sciences, Vienna Biocenter, Dr. Bohrgasse 9, 1030 Vienna, Austria; email: heribert.hirt@unvie.ac.at

Key Words programmed cell death, abiotic stress, pathogen defense s Abstract Several reactive oxygen species (ROS) are continuously produced in plants as byproducts of aerobic metabolism. Depending on the nature of the ROS species, some are highly toxic and rapidly detoxied by various cellular enzymatic and nonenzymatic mechanisms. Whereas plants are surfeited with mechanisms to combat increased ROS levels during abiotic stress conditions, in other circumstances plants appear to purposefully generate ROS as signaling molecules to control various processes including pathogen defense, programmed cell death, and stomatal behavior. This review describes the mechanisms of ROS generation and removal in plants during development and under biotic and abiotic stress conditions. New insights into the complexity and roles that ROS play in plants have come from genetic analyses of ROS detoxifying and signaling mutants. Considering recent ROS-induced genomewide expression analyses, the possible functions and mechanisms for ROS sensing and signaling in plants are compared with those in animals and yeast.

CONTENTS
INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . GENERATION OF ROS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Biotic Strategies to Generate ROS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Abiotic Strategies to Generate ROS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ROS DETOXIFICATION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Nonenzymatic ROS Scavenging Mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Enzymatic ROS Scavenging Mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . THE ROLE OF ROS IN SIGNALING . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ROS Sensing by Histidine Kinases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ROS Activation of Mitogen-Activated Protein Kinase (MAPK) Signaling Pathways . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ROS Inhibition of Protein Phosphatases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ROS Activation of Transcription Factors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1543-5008/04/0602-0373$14.00 374 374 375 377 380 380 381 382 383 383 384 384

373

374

APEL

HIRT 385 386 386 387 388 388 389

ROS AS SIGNALS FOR GENE EXPRESSION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ROS AT THE INTERFACE BETWEEN BIOTIC AND ABIOTIC STRESSES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ROS and Plant Pathogen Defense . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ROS and Abiotic Stress . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ROS and Stomata . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ROS and Roots . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . CONCLUSIONS AND OPEN QUESTIONS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

Annu. Rev. Plant Biol. 2004.55:373-399. Downloaded from arjournals.annualreviews.org by CAPES on 12/19/08. For personal use only.

INTRODUCTION
The evolution of aerobic metabolic processes such as respiration and photosynthesis unavoidably led to the production of reactive oxyen species (ROS) in mitochondria, chloroplasts, and peroxisomes. A common feature among the different ROS types is their capacity to cause oxidative damage to proteins, DNA, and lipids. These cytotoxic properties of ROS explain the evolution of complex arrays of nonenzymatic and enzymatic detoxication mechanisms in plants. Increasing evidence indicates that ROS also function as signaling molecules in plants involved in regulating development and pathogen defense responses. In this review, we rst describe the biological effects and functions of ROS in plants and then discuss open questions that need to be addressed in future research.

GENERATION OF ROS
Ground state triplet molecular oxygen is a bioradical with its two outermost valence electrons occupying separate orbitals with parallel spins. To oxidize a nonradical atom or molecule, triplet oxygen would need to react with a partner that provides a pair of electrons with parallel spins that t into its free electron orbitals. However, pairs of electrons typically have opposite spins, and thus fortunately impose a restriction on the reaction of triplet molecular oxygen with most organic molecules (18, 51). However, ground state oxygen may be converted to the much more reactive ROS forms either by energy transfer or by electron transfer reactions. The former leads to the formation of singlet oxygen, whereas the latter results in the sequential reduction to superoxide, hydrogen peroxide, and hydroxyl radical (66) (Figure 1). In plants ROS are continuously produced as byproducts of various metabolic pathways localized in different cellular compartments (37). Under physiological steady state conditions these molecules are scavenged by different antioxidative defense components that are often conned to particular compartments (3). The equilibrium between production and scavenging of ROS may be perturbed by a number of adverse environmental factors. As a result of these disturbances, intracellular levels of ROS may rapidly rise (36, 74, 102, 120). Plants also generate ROS by activating various oxidases and peroxidases that produce ROS in response to certain environmental changes (1, 13, 14, 35, 109). The rapid increase in ROS

ROS

375

Figure 1 Generation of different ROS by energy transfer or sequential univalent reduction of ground state triplet oxygen.
Annu. Rev. Plant Biol. 2004.55:373-399. Downloaded from arjournals.annualreviews.org by CAPES on 12/19/08. For personal use only.

concentration is called oxidative burst (5). External conditions that adversely affect the plants can be biotic, imposed by other organisms, or abiotic, arising from an excess or decit in the physical or chemical environment. Although plants responses to the various adverse environmental factors may show some commonalities, increases in ROS concentration triggered by either biotic or abiotic stresses are generally attributed to different mechanisms.

Biotic Strategies to Generate ROS


One of the most rapid defense reactions to pathogen attack is the so-called oxidative burst, which constitutes the production of ROS, primarily superoxide and H2O2, at the site of attempted invasion (5). Doke (1985) rst reported the oxidative burst (35), demonstrating that potato tuber tissue generated superoxide that is rapidly transformed into hydrogen peroxide following inoculation with an avirulent race of Phytopthera infestans. A virulent race of the same pathogen failed to induce O production. Subsequently, O generation has been identied in a wide range 2 2 of plant pathogen interactions involving avirulent bacteria, fungi, and viruses (71). Several different enzymes have been implicated in the generation of ROS. The NADPH-dependent oxidase system, similar to that present in mammalian neutrophils, has received the most attention. In animals the NADPH-oxidase is found in phagocytes and B lymphocytes. It catalyzes the production of superoxide by the one-electron reduction of oxygen using NADPH as the electron donor. The O generated by this enzyme serves as a starting material for the production of a 2 large variety of reactive oxidants, including oxidized halogens, free radicals, and singlet oxygen. These oxidants are used by phagocytes to kill invading microorganisms, but at the same time they may also damage surrounding cells of the host. The core of the PHagocyte OXidase comprises ve components: p40PHOX, p47 PHOX, p67PHOX, p22PHOX, and gp91PHOX. In the resting cell, three of these ve components, p40PHOX, p47PHOX, and p67PHOX, exist in the cytosol as a complex. The other two components, p22PHOX and gp91PHOX, are localized in membranes of secretory vesicles. Separating these two groups of components ensures that the oxidase remains inactive in the resting cell. When the resting cells are stimulated, the cytosolic component p47PHOX becomes heavily phosporylated and the entire

376

APEL

HIRT

cytosolic complex is recruited to the membrane, where it associates with the two membrane-bound components to assemble the active oxidase. Activation requires not only the assembly of the core components, but also the participation of two low molecular weight guanin-nucleotide-binding proteins (8). In addition to the NADPH-oxidase of phagocytes, other NADPH-oxidases also associated with plasma membranes are found in a variety of cells (8). Thus, the NADPH-oxidase activity originally described by Doke (35) may not necessarily represent a homologue of the leukocyte-specic enzyme. However, subsequent studies have provided several lines of evidence strongly suggesting a common origin for both enzymes. Antibodies raised against human p22PHOX, p47PHOX, and p67PHOX cross-reacted with plant proteins of similar sizes (32, 116), and in several plant species rboh genes (respiratory burst oxidase homologues) of p91PHOX, the catalytic subunit of the NADPH-oxidase of phagocytes, have been found (64, 119). In rice, homologues of GTP-binding proteins, required for activating the animal enzyme, have been implicated in pathogen-induced cell death (63), and the putative plant plasma membrane NADPH-oxidase reportedly produces superoxide (108). Finally, knock-out mutations of two Arabidopsis rboh genes, AtrbohD and AtrbohF, largely eliminate ROS production during disease resistance reactions of Arabidopsis to avirulent pathogens (118), thus providing direct genetic evidence that two components of a plant NADPH-oxidase are required for ROS production during plant defense responses. Homologues of the p47PHOX and p67PHOX regulators of the mammalian NADPH-oxidase were not found in the Arabidopsis genome (26), which suggests that the plant NADPH-oxidase may be regulated differently than that in mammalian macrophages. In addition to a plant-specic NADPH-oxidase, alternative mechanisms of ROS production have been proposed. Many peroxidases are localized in the apoplastic space and are ionically or covalently bound to cell wall polymers. Peroxidases can act in two different catalytic modes. In the presence of H2O2 and phenolic substrates they operate in the peroxidatic cycle and are engaged in the synthesis of lignin and other phenolic polymers. However, if the phenolic substrates are replaced by NADPH or related reduced compounds, a chain reaction starts that provides the basis for the H2O2-producing NADH-oxidase activity of peroxidases (22). Peroxidase H2O2 production is distinguished from that by the phagocyte-type NADPH-oxidase by different Km values for oxygen, different requirements for NADH and NADPH, and different sensitivities of the two enzymes to inhibitors such as cyanide, azide, and diphenyleneiodonium (DPI). Based on these differences rapid H2O2 production in some plant species triggered by pathogen attack and elicitor treatment has been attributed to the NAD(P)H-oxidase activity of apoplastic peroxidases (124). In addition to its NAD(P)H-oxidase activity that gives rise to superoxide and hydrogen peroxide, in vitro studies of horseradish peroxidase suggest another activity of this enzyme: generating hydroxyl radicals (22). Similar to the Fe2+/3+ catalyzed Haber-Weiss reaction, horseradish peroxidase can reduce hydrogen peroxide to hydroxyl radicals (22). Thus, one expects that whenever cell wallbound peroxidases

Annu. Rev. Plant Biol. 2004.55:373-399. Downloaded from arjournals.annualreviews.org by CAPES on 12/19/08. For personal use only.

ROS

377

Annu. Rev. Plant Biol. 2004.55:373-399. Downloaded from arjournals.annualreviews.org by CAPES on 12/19/08. For personal use only.

come into contact with suitable concentrations of superoxide and H2O2, originating from the oxidative cycle of peroxidase or from other sources, hydroxyl radicals should form within the cell (110). This situation prevails, for instance, when O 2 and H2O2 levels are increased in plants in response to pathogen attack, followed by a hypersensitive reaction that leads to the death of host cells. However, producing hydroxyl radicals by cell wallbound peroxidases may also be relevant for other physiological responses such as the controlled breakdown of structural polymers during rearrangement of cell walls in roots, hypocotyls, or coleoptiles (40, 94, 126). In isolated tobacco epidermal cells two distinct ROS-producing mechanisms were activated after the addition of a fungal elicitor (1). One source of ROS production was identied as a NADPH-oxidase and/or a xanthine oxidase, whereas the second activity was attributed to a peroxidase and/or amine oxidase. It is not known whether these activities appear sequentially and thus are responsible for the two phases of ROS induction by fungal or bacterial elicitors that have been measured in plant cell cultures (9). Chewing herbivorous insects mechanically wound plant tissue while feeding and induce an increase in hydrogen peroxide levels inside the plant reminiscent of the oxygen burst triggered by pathogens (12, 16). Based on the inhibition by DPI H2O2 production was linked to the NADPH-oxidase (96), thus invoking a common mechanism for the production of H2O2 during plant-pathogen interaction and the wound response. Hydrogen peroxide has been proposed to act as a second messenger for the induction of defense genes in response to wounding. However, these defense genes are different from those induced during plant-pathogen interactions. Thus, the specicity of the wound-induced defense response may not be derived from H2O2. Because the selectivity of the inhibitor used in the study (96) has been questioned (1, 39), a different source of H2O2 production cannot be ruled out.

Abiotic Strategies to Generate ROS


In plants, ROS are continuously produced predominantly in chloroplasts, mitochondria, and peroxisomes. Production and removal of ROS must be strictly controlled. However, the equilibrium between production and scavenging of ROS may be perturbed by a number of adverse abiotic stress factors such as high light, drought, low temperature, high temperature, and mechanical stress (36, 74, 102, 120).
CHLOROPLASTS HYDROGEN PEROXIDE/SUPEROXIDE Oxygen is continuously produced during light-driven photosynthetic electron transport and simultaneously removed from chloroplasts by reduction and assimilation. There are three types of oxygen-consuming processes closely associated with photosynthesis: (a) the oxygenase reaction of ribulose-1,5 bisphosphate carboxylase-oxygenase (Rubisco), (b) direct reduction of molecular oxygen by photosystem I (PSI) electron transport, and (c) chlororespiration (64).

378

APEL

HIRT

Chlororespiration describes the reduction of oxygen resulting from the presence of a respiratory chain consisting of a NAD(P)H dehydrogenase and a terminal oxidase in chloroplasts that competes with the photosynthetic electron transport chain for reducing equivalents. This process has been largely studied in microalgae (11). Only recently have components of this respiratory chain also been found in higher plants (92). It is not known to what extent this process might contribute to ROS formation in chloroplasts of higher plants although the electron transport capacity of chlororespiration is only about <1% that of photosynthesis (e.g., 36a). The two primary processes involved in formation of ROS during photosynthesis are the direct photoreduction of O2 to the superoxide radical by reduced electron transport components associated with PSI and reactions linked to the photorespiratory cycle, including Rubisco in the chloroplast and glycolate-oxidase and CAT-peroxidase reactions in the peroxisome. When plants are exposed to light intensities that exceed the capacity of CO2 assimilation, overreduction of the electron transport chain leads to inactivation of PSII and the inhibition of photosynthesis. Plants may use two strategies to protect the photosynthetic apparatus against photoinhibition: rst, the thermal dissipation of excess excitation energy in the PSII antennae (nonphotochemical quenching), and second, the ability of PSII to transfer electrons to various acceptors within the chloroplast (photochemical quenching) (98). When plants are exposed to environmental stresses and the availability of CO2 within the leaf is restricted, as may occur, for example, under drought or temperature stress, the reduction of oxygen by PSI (Mehler reaction) (79) and the photorespiratory pathway play a critical photoprotective role (Figure 2). In leaves of C3 plants the photorespiratory oxygenation of ribulose 1,5bisphosphate by Rubisco constitutes a major alternative sink for electrons, thereby sustaining partial oxidation of PSII acceptors and preventing photoinactivation of PSII when CO2 availability is restricted. Rubisco catalyzes a competitive reaction in which oxygen is favored over CO2 as a substrate as temperature increases or as the intracellular CO2 concentration declines. This oxygenation reaction leads to the release of glycolate that is translocated from chloroplasts to peroxisomes. Its subsequent oxidation is catalyzed by the glycolate oxidase accounting for the major portion of H2O2 produced during photosynthesis (Figure 2). Oxygen reduction sustains signicant levels of photosynthetic electron ux not only through its role in photorespiration but also by its direct reduction by PSI (7) (Figure 2). Superoxide radicals generated by the one-electron reduction of molecular oxygen by PSI are rapidly converted within the chloroplast to hydrogen peroxide by CuZn-superoxide dismutase (7). It has been suggested that photoreduction of O2 to water by the Mehler ascorbate peroxidase pathway in intense light may involve up to 30% of the total electron transport (7). This would suggest that O2 plays an important role as an alternative electron acceptor in photoprotection. Producing large amounts of ROS is an inevitable consequence of the photosynthetic reduction of oxygen, and plants would have to evolve efcient strategies to cope with the accumulation of these potentially toxic compounds that are integral components of oxygenic photosynthesis.

Annu. Rev. Plant Biol. 2004.55:373-399. Downloaded from arjournals.annualreviews.org by CAPES on 12/19/08. For personal use only.

ROS

379

Annu. Rev. Plant Biol. 2004.55:373-399. Downloaded from arjournals.annualreviews.org by CAPES on 12/19/08. For personal use only.

Figure 2 The principal features of photosynthetic electron transport under high light stress that lead to the production of ROS in chloroplasts and peroxisomes. Two electron sinks can be used to alleviate the negative consequences of overreduction of the photosynthetic electron chain: (a) the reduction of oxygen by PSI that generates superoxide and H2O2, and (b) the Rubisco oxygenase reaction and the photorespiratory pathway that lead to H2O2 generation within the peroxisome. Under light stress, increasing amounts of singlet oxygen are produced within PSII. Bold arrows show the main routes of electron transport. Key enzymes discussed in the text are shown in encircled numbers: 1) superoxide dismutase, 2) Rubisco, 3) glycolate oxidase, 4) catalase, and 5) ascorbate peroxidase.

Singlet Oxygen During photosynthesis singlet oxygen is continuously produced by PSII. The reaction center complex of PSII consists of cytochrome b559 and the heterodimer of the D1 and D2 proteins. The heterodimer binds the reaction centers functional prothestic groups including chlorophyll P680, pheophytin, and the quinone electron acceptors QA and QB. Excitation of the reaction center results in charge separation between P680 and pheophytin and the subsequent sequential reduction of QA and QB (10). When the redox state of the plastoquinone pool and QA and QB are overreduced because of excess light energy, charge separation cannot be completed and the oxidized P680 chlorophyll recombines with the reduced pheophytin. Under these conditions forming the triplet state of P680 is favored, leading to the generation of singlet oxygen by energy transfer. The release of singlet oxygen was rst detected in preparations of isolated PSII particles (73), but was subsequently shown to occur in vivo (41, 54). During excess light stress that leads to photoinhibition of PSII, singlet oxygen production drastically increases (53). In animals singlet oxygen may be produced metabolically during reduction of O2 catalyzed by the phagocytic NADPH-oxidase (113). A homologue of this enzyme is activated during plant-pathogen interaction. Thus, it is conceivable that singlet oxygen also takes part in defense reactions of plants. However, there is presently no data to support this proposal.

380

APEL

HIRT

Annu. Rev. Plant Biol. 2004.55:373-399. Downloaded from arjournals.annualreviews.org by CAPES on 12/19/08. For personal use only.

MITOCHONDRIA In mammalian cells mitochondria are the major source of ROS (51). However, the relative contribution of mitochondria to ROS production in green tissues is very low (103). One reason that plant mitochondria do not produce more ROS could be the presence of the alternative oxidase (AOX) that catalyzes the tetravalent reduction of O2 by ubiquinone. The AOX competes with the cytochrome bc1 complex for electrons and thus may help to reduce ROS production in mitochondria. This suggestion is supported by ndings that H2O2 induces the expression of AOX (128), and overproduction of AOX in transgenic cell lines reduces ROS production, whereas antisense cells with reduced levels of the AOX accumulate ve times more ROS than control cells (77).

ROS DETOXIFICATION
In the presence of transition metal ions hydrogen peroxide may be reduced to hydroxyl radicals by superoxide. Superoxide and hydrogen peroxide are much less reactive than OH. The main risk for a cell that produces the two former reactive oxygen intermediates may be posed by the intermediates interaction, leading to the generation of highly reactive hydroxyl radicals. Because there are no known scavengers of hydroxyl radicals, the only way to avoid oxidative damage through this radical is to control the reactions that lead to its generation. Thus, cells had to evolve sophisticated strategies to keep the concentrations of superoxide, hydrogen peroxide, and transition metals such as Fe and Cu under tight control.

Nonenzymatic ROS Scavenging Mechanisms


Nonenzymatic antioxidants include the major cellular redox buffers ascorbate and glutathione (GSH), as well as tocopherol, avonoids, alkaloids, and carotenoids. Mutants with decreased ascorbic acid levels (23) or altered GSH content (24) are hypersensitive to stress. Whereas GSH is oxidized by ROS forming oxidized glutathione (GSSG), ascorbate is oxidized to monodehydroascorbate (MDA) and dehydroascorbate (DHA). Through the ascorbate-glutathione cycle (Figure 3c), GSSG, MDA, and DHA can be reduced reforming GSH and ascorbate. In response to chilling, heat shock, pathogen attack, and drought stress, plants increase the activity of GSH biosynthetic enzymes (123, 125) and GSH levels (93). A high ratio of reduced to oxidized ascorbate and GSH is essential for ROS scavenging in cells. Reduced states of the antioxidants are maintained by glutathione reductase (GR), monodehydroascorbate reductase (MDAR), and dehydroascorbate reductase (DHAR), using NADPH as reducing power (6, 120). In addition, the overall balance among different antioxidants must be tightly controlled. The importance of this balance is evident when cells with enhanced glutathione biosynthesis in chloroplasts show oxidative stress damage, possibly due to changes of the overall redox state of chloroplasts (24). Little is known about avonoids and carotenoids in ROS detoxication in plants. However, overexpression of -carotene hydroxylase in Arabidopsis leads to increased amounts of xanthophyll in chloroplasts

ROS

381

Annu. Rev. Plant Biol. 2004.55:373-399. Downloaded from arjournals.annualreviews.org by CAPES on 12/19/08. For personal use only.

Figure 3 The principal modes of enzymatic ROS scavenging by superoxide dismutase (SOD), catalase (CAT), the ascorbate-glutathione cycle, and the glutathione peroxidase (GPX) cycle. SOD converts hydrogen superoxide into hydrogen peroxide. CAT converts hydrogen peroxide into water. Hydrogen peroxide is also converted into water by the ascorbate-glutathione cycle. The reducing agent in the rst reaction catalyzed by ascorbate peroxidase (APX) is ascorbate, which oxidizes into monodehydroascorbate (MDA). MDA reductase (MDAR) reduces MDA into ascorbate with the help of NAD(P)H. Dehydroascorbate (DHA) is produced spontaneously by MDA and can be reduced to ascorbate by DHA reductase (DHAR) with the help of GSH that is oxidized to GSSG. The cycle closes with glutathione reductase (GR) converting GSSG back into GSH with the reducing agent NAD(P)H. The GPX cycle converts hydrogen peroxide into water using reducing equivalents from GSH. Oxidized GSSG is again converted into GSH by GR and the reducing agent NAD(P)H.

and results in enhanced tolerance towards oxidative stress induced in high light (27).

Enzymatic ROS Scavenging Mechanisms


Enzymatic ROS scavenging mechanisms in plants include superoxide dismutase (SOD), ascorbate peroxidase (APX), glutathione peroxidase (GPX), and CAT (Figure 3). SODs act as the rst line of defense against ROS, dismutating superoxide to H2O2 (Figure 3a). APX, GPX, and CAT subsequently detoxify H2O2. In contrast to CAT (Figure 3b), APX requires an ascorbate and GSH regeneration system, the ascorbate-glutathione cycle (Figure 3c). Detoxiying H2O2 to

382

APEL

HIRT

Annu. Rev. Plant Biol. 2004.55:373-399. Downloaded from arjournals.annualreviews.org by CAPES on 12/19/08. For personal use only.

H2O by APX occurs by oxidation of ascorbate to MDA (equation 1 in Figure 3c), which can be regenerated by MDA reductase (MDAR) using NAD(P)H as reducing equivalents (equation 2 in Figure 3c). MDA can spontaneously dismutate into dehydroascorbate. Ascorbate regeneration is mediated by dehydroascorbate reducatase (DHAR) driven by the oxidation of GSH to GSSG (Equation 3 in Figure 3c). Finally, glutathione reductase (GR) can regenerate GSH from GSSG using NAD(P)H as a reducing agent. Like APX, GPX also detoxies H2O2 to H2O, but uses GSH directly as a reducing agent (equation 1 in Figure 3d). The GPX cycle is closed by regeneration of GSH from GSSG by GR (equation 2 in Figure 3d). Unlike most organisms, plants have multiple genes encoding SOD and APX. Different isoforms are specically targeted to chloroplasts, mitochondria, peroxisomes, as well as to the cytosol and apoplast (6). Whereas GPX is cytoslic, CAT is located mainly in peroxisomes. Specic roles for antioxidant enzymes have been explored via transgenic approaches. Overexpression of tobacco chloroplast SOD to chloroplasts did not alter tolerance toward oxidative stress, which suggests that other antioxidant mechanisms might be limiting (2). However, expression of a pea chloroplast SOD in tobacco increased resistance to methyl viologeninduced membrane damage (2). CAT is indispensable for oxidative stress tolerance because transgenic tobacco plants with suppressed CAT have enhanced ROS levels in response to both abiotic and biotic stresses (130). The extent of oxidative stress in a cell is determined by the amounts of superoxide, H2O2, and hydroxyl radicals. Therefore, the balance of SOD, APX, and CAT activities will be crucial for suppressing toxic ROS levels in a cell. Changing the balance of scavenging enzymes will induce compensatory mechanisms. For example, when CAT activity was reduced in plants, scavenging enzymes such as APX and GPX were upregulated. Unexpected effects can also occur. When compared to plants with suppressed CAT, plants lacking both APX and CAT were less sensitive to oxidative stress (106). Because photosynthetic activity of these plants was decreased, reduction in APX and CAT might result in suppression of ROS production via chloroplasts.

THE ROLE OF ROS IN SIGNALING


ROS generation in cellular compartments such as the mitochondria or chloroplasts results in changes of the nuclear transcriptome, indicating that information must be transmitted from these organelles to the nucleus, but the identity of the transmitting signal remains unknown. Three principal modes of action indicate how ROS could affect gene expression (Figure 4). ROS sensors could be activated to induce signaling cascades that ultimately impinge on gene expression. Alternatively, components of signaling pathways could be directly oxidized by ROS. Finally, ROS might change gene expression by targeting and modifying the activity of transcription factors.

ROS

383

ROS Sensing by Histidine Kinases


In prokaryotes and fungi two-component signaling systems function as redox sensors (104, 129). In prokaryotes, two-component signaling systems usually consist of a histidine kinase that senses the signal and a response regulator that functions as a transcription factor. The transmembrane sensory kinase functions through its capacity to autophosphorylate a histidine residue in response to the presence or absence of an external stimulus. The phosphoryl group is subsequently transferred from the histidine to an aspartate residue in the response regulator. The induced conformational change in the response regulator alters its DNA binding afnity and thus promotes gene expression of certain promoters. Also in budding and ssion yeast, histidine kinases of two-component signaling systems can function as sensors of oxidative stress (111). In contrast to animals, plants contain a range of two-component histidine kinases (56). Whether some of these proteins can function as ROS sensors is currently under investigation.

Annu. Rev. Plant Biol. 2004.55:373-399. Downloaded from arjournals.annualreviews.org by CAPES on 12/19/08. For personal use only.

ROS Activation of Mitogen-Activated Protein Kinase (MAPK) Signaling Pathways


Although histidine kinases are part of two-component signal transduction systems in prokaryotes that act on their own, in fungi and plants these sensors are integrated into more complex pathways. The yeast Sln1 kinase transfers its phosphoryl group via the intermediary component Ypd1 to its nal destination in the response regulator Ssk1. Stress inhibits autophosphorylation of Sln1 and hence the nonphosphorylated form of Ssk1 accumulates and activates the Hog1 mitogen-activated protein kinase (MAPK) cascade (50). Sequence analyses of the rice and Arabidopsis genomes reveal an extraordinary complexity in MAPK signaling components that comprise more than 100 MAPK, MAPKK, and MAPKKK genes in these plants. MAPK signaling modules are involved in eliciting responses to various stresses, to hormones, and during cytokinesis. H2O2 activates several MAPKs (for review see 59). In Arabidopsis, H2O2 activates the MAPKs, MPK3, and MPK6 via MAPKKK ANP1 (67). Overexpression of ANP1 in transgenic plants resulted in increased tolerance to heat shock, freezing, and salt stress (67). H2O2 also increases expression of the Arabidopsis nucleotide diphosphate (NDP) kinase 2 (85). Overexpression of AtNDPK2 reduced accumulation of H2O2 and enhanced tolerance to multiple stresses including cold, salt, and oxidative stress. The effect of NDPK2 might be mediated by the MAPKs, MPK3, and MPK6 because NDPK2 can interact and activate the MAPKs. These data suggest a scenario in which various stresses induce ROS generation that in turn activate MAPK signaling cascades. Although neither the mechanism of activation nor the downstream targets of the MAPK pathways are yet known, ROS-induced activation of MAPKs appears to be central for mediating cellular responses to multiple stresses.

384

APEL

HIRT

ROS Inhibition of Protein Phosphatases


Because H2O2 is a mild oxidant that can oxidize thiol residues, it has been speculated that H2O2 is sensed via modication of thiol groups in certain proteins. Recent work has identied human protein tyrosine phosphatase PTP1B to be modied by H2O2 at the active site cysteine (122). Inactivation of PTP1B by H2O2 is reversible and can be brought about by incubation with glutathione. A similar regulation likely occurs in plants because PTP1, an Arabidopsis PTP that can inactivate the Arabidopsis MPK6, can be inactivated by H2O2 (49). Also, phosphatases involved in abscisic acid (ABA) signaling within guard cells have been identied whose in vitro activity was modulated reversibly by H2O2 (80, 81).

Annu. Rev. Plant Biol. 2004.55:373-399. Downloaded from arjournals.annualreviews.org by CAPES on 12/19/08. For personal use only.

ROS Activation of Transcription Factors


Comparing the mechanisms for ROS-induced gene expression in prokaryotes, fungi, and plants may reveal common mechanisms (44). In E. coli, the transcription factor OxyR is of paramount importance in oxidative stress signaling (133). In budding yeast, Yap1 plays a similar role. Budding yeast mutants decient in Yap1 reveal that most ROS-induced genes depend on this transcription factor. OxyR and Yap1 are redox-sensitive transcription factors and modulate gene expression in response to oxidative stress. ROS regulates the activity of the transcription factors through covalent modication of cysteine thiol groups in OxyR and Yap1. Different types of ROS react with different cysteinyl residues and can give rise to differently modied products, possibly explaining how ROS species can induce different sets of genes via the same transcription factor (28, 29). One major difference between OxyR and Yap1 is that the yeast transcription factor is not sensing ROS directly but through the activity of Gpx3, which acts as hydroperoxidase and peroxiredoxin. The higher degree of complexity in yeast reects the increased exibility of eukaryote signaling systems. Accordingly, it is not surprising that additional regulation of redox-sensitive transcription factors was established in ssion yeast (115). Similar to yeast, plants have evolved a MAPK pathway and several protein phosphatases for ROS signaling. Although no redox-sensitive transcription factor has yet been identied in plants, it is likely that such transcription factors exist. Gene expression in response to oxidative stress seems to be coordinated via the interaction of transcription factors with specic oxidative stress-sensitive ciselements in the promoters of these genes. There is evidence that oxidative stressresponsive cis-elements exist in yeasts, animals, and plants. Work in budding and ssion yeast has shown that homologs of the mammalian ATF and AP-1 transcription factors function as key mediators of diverse stress signals binding to conserved cis-regions of stress-inducible promoters (21, 43). Microarray analysis of H2O2-induced gene expression in Arabidopsis indicates potential H2O2-responsive cis-elements in genes regulated by H2O2 (31). One of these elements, the as-1 promoter element, also has high homology with the redox-sensitive mammalian AP-1 cis-element (61). However, recent analysis of transgenic plants indicates that

ROS

385

ROS other than H2O2 activate this as-1 element (42). Further analysis will reveal whether similarity exists among plant, animal, and fungal regulatory cis-elements of ROS-responsive genes.

ROS AS SIGNALS FOR GENE EXPRESSION


Transcriptome analysis with full genome chips has revolutionized our knowledge regarding gene expression. Oxidative stress affects approximately 10% of the yeast transcriptome (19, 21, 43). Exposure of yeast cells to various stresses including H2O2 denes a large set of genes denoted as common environmental stress response (CESR). CESR-induced genes play a role in carbohydrate metabolism, ROS detoxication, protein folding and degradation, organellar function, and metabolite transport. CESR-repressed genes are involved in energy consumption and growth, RNA processing, transcription, translation, and ribosome and nucleotide biosynthesis (19, 21, 43). In plants, ROS-induced genes have been identied for receptor kinase (33), annexin (86), and peroxisome biogenesis (33). Recent approaches using cDNA proling and DNA microarrays have analyzed large-scale gene expression in response to ROS. Following exposure of Arabidopsis cells to H2O2, a total of 175 genes (i.e., 12% of the 11,000 genes on the microarray) showed changes in expression levels (31). Of the 113 induced genes, several encoded for proteins with antioxidant functions or were associated with defense responses or other stresses. Still others coded for proteins with signaling functions. Exposing a plant to sublethal doses of one stress that results in protection from lethal doses of the same stress at a later time is termed stress acclimation. Global changes in gene expression were analyzed in tobacco plants that were treated with superoxide-generating methyl viologen after pretreatment with a sublethal doses (127). Approximately 2% of the tobacco genes were altered in their expression in acclimated leaves. Genes with predicted protective or detoxifying functions and signal transduction were upregulated in acclimated leaves, implying a variety of cellular responses during acclimation tolerance. The effects of oxidative stress on the Arabidopsis mitochondrial proteome have been analyzed (114). Whereas two classes of antioxidant defense proteins, peroxiredoxins, and protein disulphide isomerase accumulated in response to oxidative stress, proteins associated with the TCA cycle were less abundant. By inhibiting H2O2 production, or facilitating its removal with scavengers such as CAT, genes encoding APX, pathogenesis-related (PR) proteins, glutathione Stransferase (GST), and phenylalanine ammonia-lyase (PAL) were identied (34, 62, 70). An alternative approach to study the effects of oxidative stress on the transcriptome is to induce oxidative stress by reducing antioxidant activity. CAT and ascorbate peroxidase antisense lines show elevated expression of SOD and GR (106). In contrast, MDAR, a key enzyme for the regeneration of ascorbate, was upregulated in plants with experimentally reduced CAT and ascorbate peroxidase

Annu. Rev. Plant Biol. 2004.55:373-399. Downloaded from arjournals.annualreviews.org by CAPES on 12/19/08. For personal use only.

386

APEL

HIRT

levels. An increase in expression of ROS detoxifying enzymes is compatible with compensatory mechanisms induced by oxidative stress. When tobacco plants decient in CAT were grown in high-intensity light, they increased ROS production and PR protein levels, and showed enhanced disease resistance (20).

ROS AT THE INTERFACE BETWEEN BIOTIC AND ABIOTIC STRESSES


Annu. Rev. Plant Biol. 2004.55:373-399. Downloaded from arjournals.annualreviews.org by CAPES on 12/19/08. For personal use only.

ROS and Plant Pathogen Defense


ROS play a central role in plant pathogen defense. During this response, ROS are produced by plant cells via the enhanced enzymatic activity of plasma-membranebound NADPH-oxidases, cell wallbound peroxidases and amine oxidases in the apoplast (Figure 5a) (47, 52). Under these conditions, up to 15 M H2O2 can be produced either directly or as a result of superoxide dismutation. In contrast to superoxide, H2O2 can diffuse into cells and activate many of the plant defenses, including PCD (programmed cell death) (26). During plant pathogen reactions, the activity and levels of the ROS detoxifying enzymes APX and CAT are suppressed by salicylic acid (SA) and nitrous oxide (NO) (65). Because during the plant pathogen defense reponse the plant simultaneously produces more ROS while decreasing its ROS scavenging capacities, accumulation of ROS and activation of PCD occurs. The suppression of ROS detoxifying mechanisms is crucial for the onset of PCD. ROS production at the apoplast alone without suppression of ROS detoxication does not result in the induction of PCD (30, 83). These data indicate an absolute requirement for the coordinated production of ROS and downregulation of ROS scavenging mechanisms. Induction of PCD potentially limits the spread of disease from the infection point. During incompatible reactions, when a pathogen is detected as an enemy and defense responses including PCD are induced, H2O2 production occurs in a biphasic manner. The initial and very rapid accumulation of H2O2 is followed by a second and prolonged burst of H2O2 production. During compatible interactions, when a pathogen overcomes the defense lines and systemically infects the host plant, only the rst peak of H2O2 accumulation occurs (9). It is not yet known whether these two distinct bursts arise from the same or different sources. H2O2 generation occurs both locally and systemically in response to wounding (97). Recent work shows that H2O2 functions as a second messenger mediating the systemic expression of various defense-related genes in tomato plants (96). Previously, it was found that the oxidative burst in pathogen challenged Arabidopsis leaves activates a secondary systemic burst in distal parts of the plant, leading to systemic immunity via the expression of defense-related genes (4). It is possible that H2O2 is not the primary signal that is transmitted, and interactions with other signaling intermediates such as SA could also be involved. Although the oxidative burst is a primary response to pathogen challenge that leads to PCD (15), and H2O2 induces PCD in various systems (34, 70, 112),

ROS

387

Annu. Rev. Plant Biol. 2004.55:373-399. Downloaded from arjournals.annualreviews.org by CAPES on 12/19/08. For personal use only.

in some cases H2O2 is not required for PCD induction (46, 57). Studies show that a threshold exposure time of cells to H2O2 is required, during which period transcription and translation are necessary (34, 112). Pharmacological data indicate that removal of ROS during pathogen or elicitor challenge reduces PCD (34, 70). A recent genetic analysis conrms these data by showing that Arabidopsis knock-out lines lacking functional rboh genes (i.e., respiratory burst oxidase homolog genes) display reduced ROS generation and PCD following bacterial challenge (118). In line with this concept, tobacco plants with reduced CAT or ascorbate peroxidase expression show increased PCD to low doses of bacteria (83). It is not clear yet to what extent PCD in plants and animals share similar mechanisms. Expression of the animal cell death suppressor genes Bcl-xl and Ced-9 in tobacco plants suppresses oxidative stress-induced cell death (82). In animals, mitochondria play a primary role in ROS production and in triggering programmed cell death. Mitochondrial ROS production has also been implicated in eliciting ROS-induced cell death in plants (76, 117). However, as mentioned earlier, in plants the majority of ROS is produced in chloroplasts and peroxisomes. Whether or not these differences necessitated the evolution of different mechanisms of PCD in plants and animals is not known yet.

ROS and Abiotic Stress


PCD occurs not only as a result of the oxidative burst following pathogen challenge, but also following exposure to abiotic stresses. For example, the ozone-induced oxidative burst results in a cell death process similar to the hypersensitivity response (HR) during plant-pathogen interactions (131). However, the role that ROS play during abiotic stresses appears to be opposite to the role that ROS play during pathogen defense. Upon abiotic stresses, ROS scavenging enzymes are induced to decrease the concentration of toxic intracellular ROS levels (Figure 5b). The differences in the function of ROS between biotic and abiotic stresses might arise from the action of hormones and cross-talk between different signaling pathways or from differences in the locations where ROS are produced and/or accumulate during different stresses. These considerations raise the question of how plants can regulate ROS production and scavenging mechansisms when they are exposed simultaneously to pathogen attack and abiotic stress. Evidence for the signicance of such conicting situations comes from experiments with tobacco plants showing reduced PCD after exposure to oxidative stress (83). The oxidative stress pretreatment resulted in increased levels of ROS scavenging enzymes, thereby abrogating the plants ability to build up sufcient ROS for inducing PCD. In accordance with this model, CAT overproducing plants have decreased resistance to pathogen infection (101), wounding (97), and high light treatment (88). It was suggested that ROS act in conjunction with compound(s) that travel systemically and have the capacity to activate ROS production in distant parts of the plant (97). It is still debated whether ROS can travel long distances in the plant because most ROS are highly reactive and are detoxied immediately by the scavenging systems of the

388

APEL

HIRT

apoplast. Future studies using plants with altered levels of ROS scavenging and/or ROS producing mechanisms might resolve this question.

ROS and Stomata


Recent work has shown that ROS are essential signals mediating ABA-induced stomatal closure. The phytohormone ABA accumulates in response to water stress and induces a range of stress adaptation responses including stomatal closure. Earlier work has shown that H2O2 induces stomatal closure (78) and that guard cells synthesize ROS in response to elicitor challenge (1, 68). H2O2 is an endogenous component of ABA signaling in Arabidopsis guard cells. ABA-stimulated ROS accumulation induced stomatal closure via activation of plasma membrane calcium channels (99). ABA-induced ROS synthesis also occurs in Vicia faba (132), but ROS production occurs at the plasma membrane and in the chloroplast. This study indicates the complexity of ROS signaling in this system. Various Arabidopsis mutants have been used to dissect ABA and ROS signaling in guard cells. In the gca2 mutant, ABA increased ROS production, but H2O2-induced calcium channel activation and stomatal closure were absent in the mutant (99). Protein phosphorylation is also involved in guard cell signaling, as shown by analysis of the ABA-insensitive abi1 and abi2 mutants. ABI1 and ABI2 encode protein phosphatase 2C enzymes that are both involved in stomatal closing. Using the abi1 and abi2 point mutants with strongly reduced phosphatase activities, it was shown that ABA is unable to generate ROS in abi1 mutants but ABA still induces ROS production in abi2 mutants (89). These data indicate that ABI1 may act upstream and ABI2 downstream of ROS signaling. A recently identied protein kinase functions between ABA perception and ROS signaling (90). Ost1 was identied as an ABA insensitive mutant. OST1 kinase is activated by ABA in guard cell protoplasts of wild-type but not of ost1 plants. ABA-induced ROS production was absent in ost1 plants, although ost1 stomata still closed in response to H2O2. The notion that OST1 regulates ROS production directly via the NADPH-oxidase is an attractive hypothesis that remains to be validated experimentally. As shown by the recent ndings that NO can also mediate ABA-induced stomatal closure (91), guard cell behavior is probably not solely regulated by ROS.

Annu. Rev. Plant Biol. 2004.55:373-399. Downloaded from arjournals.annualreviews.org by CAPES on 12/19/08. For personal use only.

ROS and Roots


A new role for H2O2 in auxin signaling and gravitropism in maize roots was revealed recently (60). Gravity and asymmetric auxin application induced ROS generation, and asymmetric application of H2O2 promoted gravitropism. An intracellular source of ROS was suggested because CAT application had no effect on gravitropism. The upregulation of oxidative stress-related genes during Arabidopsis gravitropism might indicate a wider role of ROS in this biological process (86).

ROS

389

CONCLUSIONS AND OPEN QUESTIONS


Although there has been rapid progress in recent years, there are still many uncertainties and gaps in our understanding of how ROS affect the stress response of plants. Here we focus on some of the more important questions that need to be addressed in the future. Generally, ROS have been proposed to affect stress responses in two different ways. ROS react with a large variety of biomolecules, and may thus cause irreversible damage that can lead to tissue necrosis and may ultimately kill the plants (45, 105). On the other hand, ROS inuence the expression of a number of genes and signal transduction pathways. These latter observations suggest that cells have evolved strategies to utilize ROS as environmental indicators and biological signals that activate and control various genetic stress response programs (25). This interpretation is based on the unstated assumption that a given ROS may interact selectively with a target molecule that perceives the increase of ROS concentration and translates this information into signals that direct the plants responses to stress. ROS would be ideally suited to act as such signaling molecules. ROS are small and can diffuse short distances, and there are several mechanisms for ROS production, some of which are rapid and controllable, and there are numerous mechanisms for rapid removal of ROS. There are at least three major possibilites of how ROS could act as biological signals in plants. (a) ROS could act as a second messenger and modulate the activity of a specic target molecule involved in signaling or transcription, as described above. (b) Many changes in gene expression that have been attributed to a signaling role of ROS could also result from their cytotoxicity. The toxicity of ROS has often been monitored by measuring lipid peroxidation. Polyunsaturated fatty acids within the lipids are a preferred target of ROS attack. Several of their oxygenation products are biologically active and may change the expression of specic genes (48). Thus, a given ROS may generate nonenzymatically a wide range of oxidation products, some of which may disseminate within the cell and act as a second messenger that triggers multiple stress responses. Whether or not these effects can be ascribed to a signaling role of ROS depends on how signaling is dened. (c) Finally, ROS may trigger stress responses in plants by modulating gene expression in a more indirect way. For instance, during detoxication of ROS in chloroplasts, large amounts of reductants such as ascorbic acid and glutathione are oxidized and shift the redox balance to a more oxidized state. For example, changes in the redox status of chloroplasts during a light-dark cycle are known to modulate the activity of many enzymes (38) and to inuence the transcription of a variety of genes (38, 91). Redox changes in plants reverse the activity of the NPR1 protein, an essential regulator of plant systemic acquired resistance against pathogens (87). These few examples of how variable the effect of ROS on gene expression may be emphasize the importance of rst identifying more precisely the targets and the modes of action of ROS during gene activation in plants under oxidative stress before a signaling role of a given ROS can be rmly established.

Annu. Rev. Plant Biol. 2004.55:373-399. Downloaded from arjournals.annualreviews.org by CAPES on 12/19/08. For personal use only.

390

APEL

HIRT

Depending on the character of the environmental stress, plants differentially enhance the release of ROS that are either chemically distinct or are generated within different cellular compartments (36, 91). For instance, during an incompatible plant-pathogen interaction, superoxide anions are produced enzymatically outside the cell and are rapidly converted to hydrogen peroxide that can cross the plasma membrane. The same ROS are also produced in chloroplasts exposed to high light stress, albeit by a different mechanism. The stress reactions of plants induced by pathogens differ from those induced by high light intensities (17, 52). If ROS act as signals that evoke these different stress responses, their biological activities should exhibit a high degree of selectivity and specicity that could be derived from their chemical identity and/or the intracellular locations where they were generated. Most of what is known about oxidant-induced signaling was found in experiments using hydrogen peroxide as an oxidant. In some of these experiments hydrogen peroxide was added to cell cultures or plants either directly (31) or indirectly by using H2O2-generating enzymes (4). In both cases the intracellular distribution of H2O2 is difcult to control and the physiological signicance of changes in gene expression that are induced under these conditions cannot easily be assessed. A more controlled release of H2O2 that is conned to specic compartments has become possible either by spraying plants with paraquat, a herbicide that under light generates H2O2 via superoxide predominantly within chloroplasts (79), or by modulating the expression of scavengers that are conned to different cellular compartments in transgenic plants (107, 121, 130). If one accepts the notion that the chemical identity of a given ROS and the cellular topography of its release contribute to the specicity of a stress response, it is hard to imagine how H2O2 alone could act as a signal that triggers such specic responses unless differences in the intracellular distribution of H2O2-responsive targets mediate such specicities of stress responses. As mentioned above it is also possible that hydroxyl radicals derived from hydrogen peroxide are the primary triggers of some stress responses in plants. In contrast to hydrogen peroxide, hydroxyl radicals are very unstable and restricted to subcompartmental regions. They may give rise to biologically active oxygenation products that could preserve some of the topographical information that may be necessary to ensure specicity of stress responses. Measuring global changes of gene expression in Arabidopsis cell cultures following treatment with hydrogen peroxide reveal major changes in gene expression (31). The surprisingly large number of genes that respond to an increase in hydrogen peroxide concentration would be in line with the proposed role of H2O2 as a ubiquitous signal for oxidative stress. However, the number of H2O2-responsive genes in cell culture was in marked contrast to that found in whole plants treated with low concentrations of paraquat (95). In the absence of visible necrotic lesions, very few genes were initially upregulated, some of which are intimately involved in the detoxication of H2O2, like ascorbate peroxidases I and II or ferritin I (62,

Annu. Rev. Plant Biol. 2004.55:373-399. Downloaded from arjournals.annualreviews.org by CAPES on 12/19/08. For personal use only.

ROS

391

100). These genes were different from those activated by singlet oxygen, which suggests that chemical differences between the two ROS might have contributed to the selectivity of the induced stress responses (95). This apparent selectivity was lost when plants were exposed to higher paraquat concentrations that led to visible lesion formation. Under these stress conditions the primary cause for a given stress response may be difcult to separate from secondary effects that are superimposed. This problem of discerning cause and effect has been highlighted in studies of tetrapyrrole biosynthesis and catabolism, in which several enzymatic steps were blocked experimentally (55, 72, 75, 84). A broad range of stress-related reactions were activated in these plants during illumination that have been attributed to the photodynamic activity of free tetrapyrrole intermediates that accumulated constitutively in these plants. However, many aspects of the responses are reminescent of those triggered by the release of superoxide and/or hydrogen peroxide during an incompatible plant-pathogen interaction (84). Either singlet oxygen, superoxide, and hydrogen peroxide may replace each other in triggering pathogen defense reactions, or the constitutive accumulation of photodynamically active tetrapyrrole intermediates and catabolites throughout the entire life cycle of these genetically modied plants may lead to photooxidative damage and injury that may promote a multifactorial induction of several overlapping secondary effects, some of which may mimic responses to pathogens. This latter interpretation agrees with in vivo measurements of ROS production in leaves under photooxidative stress, showing that singlet oxygen, superoxide, and hydrogen peroxide were produced simultaneously in the same leaf (41). If chemical specicity of a given ROS determines the specicity of stress responses, it would be difcult under these conditions to attribute a particular stress response to a well-dened ROS. So far very few case studies suggest a selective signaling effect of a given ROS (58, 69, 95). This may be due partially to the fact that most of the previous studies have focused on analyzing the signaling role of hydrogen peroxide andto a lesser extentsuperoxide radicals, whereas other ROS such as hydroxyl radicals and singlet oxygen have been largely ignored. As described earlier, there are several lines of evidence suggesting that hydroxyl radicals may not only be a noxious side product of oxygen metabolism but may play a more signicant role not only during oxidative stress but also during extension growth of roots, coleoptiles, and hypocotyls, or during seed germination. Singlet oxygen induces a specic set of stress responses (95). Its biological activity exhibits a high degree of selectivity that is derived from the chemical identity of this ROS and/or the intracellular location at which it is generated. These and other studies indicate that the biological activities of ROS may signicantly differ from each other. Hence, ndings from experiments with hydrogen peroxide as a model oxidant should not be inappropriately generalized and taken as a xed template for signaling events induced by all oxidants or by oxidative stress generated in a different way (66).

Annu. Rev. Plant Biol. 2004.55:373-399. Downloaded from arjournals.annualreviews.org by CAPES on 12/19/08. For personal use only.

392

APEL

HIRT

ACKNOWLEDGMENTS This work was supported from grants from the Austrian and Swiss Science Foundations and the European Union.
The Annual Review of Plant Biology is online at http://plant.annualreviews.org

LITERATURE CITED
1. Allan AC, Fluhr R. 1997. Two distinct sources of elicited reactive oxygen species in tobacco epidermal cells. Plant Cell. 9:155972 2. Allen RD. 1995. Dissection of oxidative stress tolerance using transgenic plants. Plant Physiol. 107:104954 3. Alscher RG, Donahue JH, Cramer CL. 1997. Reactive oxygen species and antioxidants: relationships in green cells. Physiol. Plantarum. 100:22433 4. Alvarez ME, Pennell RI, Meijer P-J, Ishikawa A, Dixon RA, Lamb C. 1998. Reactive oxygen intermediates mediate a systemic signal network in the establishment of plant immunity. Cell 92:77384 5. Apostol I, Heinstein PF, Low PS. 1989. Rapid stimulation of an oxidative burst during elicidation of cultured plant cells. Role in defense and signal transduction. Plant Physiol. 90:10616 6. Asada K, Takahashi M. 1987. Production and scavenging of active oxygen in photosynthesis. In Photoinhibition, eds. DJ Kyle, CB Osborne, CJ Arntzen, pp. 227 87, Amsterdam: Elsevier 7. Asada K. 1999. The waterwater cycle in chloroplasts: scavenging of active oxygens and dissipation of excess photons. Annu. Rev. Plant Physiol. Plant Mol. Biol. 50:60139 8. Babior BM. 1999. NADPH oxidase: an update. Blood 93:146476 9. Baker MA, Orlandi EW. 1995. Active oxygen in plant pathogenesis. Annu. Rev. Phytopathol. 33:299321 10. Barber J. 1998. Photosystem two. Biochim. Biophys. Acta. 1365:26977 11. Bennoun P. 1982. Evidence for a respiratory chain in the chloroplast. Proc. Natl. Acad. Sci. USA 79:435256 Bi JL, Felton GW. 1995. Foliar oxidative stress and insect herbivory: primary compounds, secondary metabolites and reactive oxygen species as components of induced resistance. J. Chem. Ecol. 151130 Bolwell GP, Bindschedler LV, Blee KA, Butt VS, Davies DR, et al. 2002. The apoplastic oxidative burst in response to biotic stress in plants: a three-component system. J. Exp. Bot. 53:136776 Bolwell GP, Davies DR, Gerrish C, Auh CK, Murphy TM. 1998. Comparative biochemistry of the oxidative burst produced by rose and french bean cells reveals two distinct mechanisms. Plant Physiol. 116:137985 Bolwell GP. 1999. Role of active oxygen species and NO in plant defense responses. Curr. Opin. Plant Biol. 2:28794 Bradley DJ, Kjellbom P, Lamb CJ. 1992. Elicitor- and wound-induced oxidative cross-linking of a proline-rich plant cellwall protein: A novel, rapid defense response. Cell 70:2130 Bray EA, Bailey-Serres J, Weretilnyk E. 2000. Responses to abiotic stresses. In Biochemistry and Molecular Biology of Plants, eds. BB Buchanan, W Gruissem and RL Jones, pp. 1158203. Rockville, MD: Am. Soc. Plant Physiol. Cadenas E. 1989. Biochemistry of oxygen toxicity. Annu. Rev. Biochem. 58:79110 Causton HC, Ren B, Koh SS, Harbison CT, Kanin E, et al. 2001. Remodeling of yeast genome expression in response to environmental changes. Mol. Biol. Cell 12:32337
Annu. Rev. Plant Biol. 2004.55:373-399. Downloaded from arjournals.annualreviews.org by CAPES on 12/19/08. For personal use only.

12.

13.

14.

15.

16.

17.

18. 19.

ROS 20. Chamnongpol S, Willekens H, Moeder W, Langebartels C, Sandermann H Jr, et al. 1998. Defense activation and enhanced pathogen tolerance induced by H2O2 in transgenic tobacco. Proc. Natl. Acad. Sci. USA 95:581823 21. Chen D, Toone WM, Mata J, Lyne R, Burns G, et al. 2003. Global transcriptional responses of ssion yeast to environmental stress. Mol. Biol. Cell 14:214 29 22. Chen SX, Schopfer P. 1999. Hydroxylradical production in physiological reactions. A novel function of peroxidase. Eur. J. Biochem. 260:72635 23. Conklin PL, Williams EH, Last RL. 1996. Environmental stress sensitivity of an ascorbic acid-decient Arabidopsis mutant. Proc. Natl. Acad. Sci. USA 93:9970 74 24. Creissen G, Firmin J, Fryer M, Kular B, Leyland N, et al. 1999. Elevated glutathione biosynthetic capacity in the chloroplasts of transgenic tobacco plants paradoxically causes increased oxidative stress. Plant Cell 11:127792 25. Dalton TP, Shertzer HG, Puga A. 1999. Regulation of gene expression by reactive oxygen. Annu. Rev. Pharmacol. Toxicol. 39:67101 26. Dangl JL, Jones JDG. 2001. Plant pathogens and integrated defense responses to infection. Nature 411:82633 27. Davison PA, Hunter CN, Horton P. 2002. Overexpression of -carotene hydroxylase enhances stress tolerance in Arabidopsis. Nature 418:2036 28. Delaunay A, Isnard A-D, Toledano MB. 2000. H2O2 sensing through oxidation of the Yap1 transcription factor. EMBO J. 19:515766 29. Delaunay A, Pieger D, Barrault MB, Vinh J, Toledano MB. 2002. A thiol peroxidase is an H2O2 receptor and redox-transducer in gene activation. Cell 111:47181 30. Delledonne M, Zeier J, Marocco A, Lamb C. 2001. Signal interactions between ni-

393

31.

32.

33.

34.

35.

36.

36a.

37.

38.

tric oxide and reactive oxygen intermediates in the plant hypersensitive disease resistance response. Proc. Natl. Acad. Sci. USA 98:1345459 Desikan R, A-H Mackerness S, Hancock JT, Neill SJ. 2001. Regulation of the Arabidopsis transcriptome by oxidative stress. Plant Physiol. 127:15972 Desikan R, Hancock JT, Coffey MJ, Neill SJ. 1996. Generation of active oxygen in elicited cells of Arabidopsis thaliana is mediated by a NADPH oxidase-like enzyme. FEBS Lett. 382:21317 Desikan R, Neill SJ, Hancock JT. 2000. Hydrogen peroxide-induced gene expression in Arabidopsis thaliana. Free Rad. Biol. Med. 28:77378 Desikan R, Reynolds A, Hancock JT, Neill SJ. 1998. Harpin and hydrogen peroxide both initiate programmed cell death but have differential effects on gene expression in Arabidopsis suspension cultures. Biochem. J. 330:11520 Doke N. 1985. NADPH-dependent O2generation in membrane fractions isolated from wounded potato tubers inoculated with Phytophtora infestans. Physiol. Plant Pathol. 27:31122 Elstner EF. 1991. Mechanisms of oxygen activation in different compartments of plant cells. In Active Oxygen/Oxidative Stress in Plant Metabolis, eds. EJ Pelland, KL Steffen. pp.1325. Rockville, MD: Am. Soc. Plant Physiol. Feild TS, Nedbal L, Ort DR. 1998. Nonphotochemical reduction of the plastoquinone pool in sunower leaves originates from chlororespiration. Plant Physiol. 116:120918 Foyer CH, Harbinson JC. 1994. Oxygen metabolism and the regulation of photosynthetic electron transport. In Causes of Photooxidative Stress and Amelioration of Defense Systems in Plant, eds. CH Foyer, PM Mullineaux. pp. 142. Boca Raton, Fla.: CRC Foyer CH, Noctor G. 2000. Oxygen processing in photosynthesis: regulation

Annu. Rev. Plant Biol. 2004.55:373-399. Downloaded from arjournals.annualreviews.org by CAPES on 12/19/08. For personal use only.

394

APEL

HIRT trol of protein tyrosine phosphatases and mitogen-activated protein kinases in plants. Plant Physiol. 132:114952 Gustin MC, Albertyn J, Alexander M, Davenport K. 1998. MAP kinase pathways in the yeast Saccharomyces cerevisiae. Microbiol. Mol. Biol. Rev. 62:1264300 Halliwell B, Gutteridge JMC. 1989. Free Radicals in Biology and Medicine, 2nd ed., Oxford: Clarendon Hammond-Kosack K, Jones JDG. 2000. Responses to plant pathogens. In Biochemistry and Molecular Biology of Plants, eds. BB Buchanan, W Gruissem, RL Jones. pp. 1102156. Rockville, MD: Am. Soc. Plant Physiol. Hideg E, Barta C, Kalai T, Vass I, Hideg K, Asada K. 2002. Detection of singlet oxygen and superoxide with uorescent sensors in leaves under stress by photo inhibition or UV radiation. Plant Cell Physiol. 43:115464 Hideg E, Kalai T, Hideg K, Vass I. 1998. Photoinhibition of photosynthesis in vivo results in singlet oxygen production. Detection via nitroxide-induced uorescence quenching in broad bean leaves. Biochemistry 37:1140511 Hu G, Yalpani N, Briggs SP, Johal GS. 1998. A porphyrin pathway impairment is responsible for the phenotype of a dominant disease lesion mimic mutant of maize. Plant Cell 10:1095105 Hwang I, Chen H-C, Sheen J. 2002. Twocomponent signal transduction pathways in Arabidopsis. Plant Physiol. 129:500 15 Ichinose Y, Andi S, Doi R, Tanaka R, Taguchi F, et al. 2001. Generation of hydrogen peroxide is not required for harpin-induced apoptotic cell death in tobacco BY-2 cell suspension cultures. Plant Physiol. Biochem. 39:77176 Jabs T, Dietrich RA, Dangl JL. 1996. Initiation of runaway cell death in an Arabidopsis mutant by extracellular superoxide. Science 273:185356

39.

40.

41.

42.

43.

44. 45.

46.

47.

48.

49.

and signaling. New Phytol. 146:359 88 Frahry G, Schopfer P. 1998. Inhibition of O2-reducing activity of horseradish peroxidase by diphenyleneiodonium. Phytochemistry 48:22327 Frahry G, Schopfer P. 2001. NADHstimulated, cyanide-resistant superoxide production in maize coleoptiles analyzed with a tetrazolium-based assay. Planta 212:17583 Fryer MJ, Oxborough K, Mullineaux PM, Baker NR. 2002. Imaging of photooxidative stress responses in leaves. J. Exp. Bot. 53:124954 Garreton V, Carpinelli J, Jordana X, Holuigue L. 2002. The as-1 promoter element is an oxidative stress-responsive element and salicylic acid activates it via oxidative species. Plant Physiol. 130:1516 26 Gasch A, Spellman P, Kao C, Harel O, Eisen M, et al. 2000. Genomic expression programs in the response of yeast cells to environmental changes. Mol. Biol. Cell 11:424157 Georgiou G. 2002. How to ip the (redox) switch. Cell 111:60710 Girotti AW. 2001. Photosensitized oxidation of membrane lipids: reaction pathways, cytotoxic effects and cytoprotective mechanisms. J. Photochem. Photobiol. 63:10313 Glazener JA, Orlandi EW, Baker CJ. 1996. The active oxygen response of cell suspensions to incompatible bacteria is not sufcient to cause hypersensitive cell death. Plant Physiol. 110:75963 Grant JJ, Loake GJ. 2000. Role of reactive oxygen intermediates and cognate redox signaling in disease resistance. Plant Physiol. 124:2129 Grether-Beck S, Bonizzi G, SchmittBrenden H, Felsner I, Timmer A, et al. 2000. Non-enzymatic triggering of the ceramide signaling cascade by solar UVA radiation. EMBO J. 19:5793800 Gupta R, Luan S. 2003. Redox con-

50.

51.

Annu. Rev. Plant Biol. 2004.55:373-399. Downloaded from arjournals.annualreviews.org by CAPES on 12/19/08. For personal use only.

52.

53.

54.

55.

56.

57.

58.

ROS 59. Jonak C, Okresz L, B gre L, Hirt H. 2002. o Complexity, cross talk and integration of plant MAP kinase signaling. Curr. Opin. Plant Biol. 5:41524 60. Joo JH, Bae YS, Lee JS. 2001. Role of auxin-induced reactive oxygen species in root gravitropism. Plant Physiol. 126:105560 61. Karin M, Liu Z, Zandi E. 1997. AP-1 function and regulation. Curr. Opin. Cell Biol. 9:24046 62. Karpinski S, Reynolds H, Karpinska B, Wingsle G, Creissen J, Mullineaux PC. 1999. Systemic signaling and acclimation in response to excess excitation energy in Arabidopsis. Science 284:65457 63. Kawasaki T, Henmi K, Ono E, Hatakeyama S, Iwano M, et al. 1999. The small GTP-binding protein Rac is a regulator of cell death in plants. Proc. Natl. Acad. Sci. USA 96:1092226 64. Keller T, Danude HG, Werner D, Doerner P, Dixon RA, Lamb C. 1998. A plant homolog of the neutrophil NADPH oxidase gp91phox subunit gene encodes a plasma membrane protein with Ca2+ binding motifs. Plant Cell. 10:25566 65. Klessig DF, Durner J, Noad R, Navarre DA, Wendehenne D, et al. 2000. Nitric oxide and salicylic acid signaling in plant defense. Proc. Natl. Acad. Sci. USA 97:884955 66. Klotz L-O. 2002. Oxidant-induced signaling: effects of peroxynitrite and singlet oxygen. Biol. Chem. 383:44356 67. Kovtun Y, Chiu W-L, Tena G, Sheen J. 2000. Functional analysis of oxidative stress-activated mitogen-activated protein kinase cascade in plants. Proc. Natl. Acad. Sci. USA 97:294045 68. Lee S, Choi H, Suh S, Doo I-S, Oh K-Y, et al. 1999. Oligogalacturonic acid and chitosan reduce stomatal aperture by inducing the evolution of reactive oxygen species from guard cells of tomato and Commelina communis. Plant Physiol. 121:14752 69. Leisinger U, Rufenacht K, Fischer B, Pe-

395

70.

71.

72.

73.

74.

75.

76.

77.

78.

saro M, Spengler A, et al. 2001. The glutathione peroxidase homologous gene from Chlamydomonas reinhardtii is transcriptionally up-regulated by singlet oxygen. Plant Mol. Biol. 46:395408 Levine A, Tenhaken R, Dixon R, Lamb C. 1994. H2O2 from the oxidative burst orchestrates the plant hypersensitive disease resistance response. Cell 79:58393 Low PS, Merida JR. 1996. The oxidative burst in plant defense: function and signal transduction. Physiol. Plant. 96:53342 Mach JM, Castillo AR, Hoogstraten R, Greenberg JT. 2001. The Arabidopsisaccelerated cell death gene ACD2 encodes red chlorophyll catabolite reductase and suppresses the spread of disease symptoms. Proc. Natl. Acad. Sci. USA 98:77176 Macpherson AN, Telfer A, Barber J, Truscott TG. 1993. Direct detection of singlet oxygen from isolated photosystem II reaction centres. Biochim. Biophys. Acta 1143:3019 Malan C, Gregling MM, Gressel J. 1990. Correlation between CuZn superoxide dismutase and glutathione reductase and environmental and xenobiotic stress tolerance in maize inbreds. Plant Sci. 69:157 66 Matringe M, Camadro JM, Labbe P, Scalla R. 1989. Protoporphyrinogen oxidase as a molecular target for diphenylether herbicides. Biochem. J. 260:23135 Maxwell DP, Nickels R, McIntosh L. 2002. Evidence of mitochondrial involvement in the transduction of signals required for the induction of genes associated with pathogen attack and senescence. Plant J. 29:26979 Maxwell DP, Wang Y, McIntosh L. 1999. The alternative oxidase lowers mitochondrial reactive oxygen production in plant cells. Proc. Natl. Acad. Sci. USA 96:8271 76 McAinsh MR, Clayton H, Manseld TA, Hetherington AM. 1996. Changes in stomatal behaviour and guard cell

Annu. Rev. Plant Biol. 2004.55:373-399. Downloaded from arjournals.annualreviews.org by CAPES on 12/19/08. For personal use only.

396

APEL

HIRT 87. Mou Z, Fan W, Dong X. 2003. Inducers of plant systemic acquired resistance regulate NPR1 function through redox changes. Cell 113:93544 88. Mullineaux P, Karpinski S. 2002. Signal transduction in response to excess light: getting out of the chloroplast. Curr. Opin. Plant Biol. 5:4348 89. Murata Y, Pei Z-M, Mori IC, Schroeder J. 2001. Abscisic acid activation of plasma membrane Ca2+ channels in guard cells requires cytosolic NAD(P)H and is differentially disrupted upstream and downstream of reactive oxygen species production in abi1-1 and abi2-1 protein phosphatase 2C mutants. Plant Cell. 13:251323 90. Mustilli A-C, Merlot S, Vavasseur A, Fenzi F, Giraudat J. 2002. Arabidopsis OST1 protein kinase mediates the regulation of stomatal aperture by abscisic acid and acts upstream of reactive oxygen species production. Plant Cell 14:3089 99 91. Neill N, Desikan R, Hancock J. 2002. Hydrogen peroxide signaling. Curr. Opin. Plant Biol. 5:38895 92. Nixon PJ. 2000. Chlororespiration. Philos. Trans. R. Soc. London B 355:154147 93. Noctor G, Gomez L, Vanacker H, Foyer CH. 2002. Interactions between biosynthesis, compartmentation and transport in the control of glutathione homeostasis and signaling. J. Exp. Bot. 53:1283304 94. Ogawa K, Kanematsu S, Asada K. 1997. Generation of superoxide anion and localization of CuZn-superoxide dismutase in the vascular tissue of spinach hypocotyls: their association with lignication. Plant Cell Physiol. 38:111826 95. op den Camp RG, Przybyla D, Ochsenbein C, Laloi C, Kim C, et al. 2003. Rapid induction of distinct stress responses following the release of singlet oxygen in Arabidopsis thaliana. Plant Cell 15:2320 32 96. Orozco-Cardenas ML, Narvaez-Vasquez J, Ryan CA. 2001. Hydrogen peroxide

79.

80.

81.

82.

83.

84.

85.

86.

cytosolic free calcium in response to oxidative stress. Plant Physiol. 111:103142 Mehler AH. 1951. Studies on reactions of illuminated chloroplasts. II Stimulation and inhibition of the reaction with molecular oxygen. Arch. Biochem. Biophys. 33:33951 Meinhard M, Grill E. 2001. Hydrogen peroxide is a regulator of ABI1, a protein phosphatase 2C from Arabidopsis. FEBS Lett. 508:44346 Meinhard M, Rodriguez PL, Grill E. 2002. The sensitivity of ABI2 to hydrogen peroxide links the abscisic acid-response regulator to redox signaling. Planta 214:775 82 Mitsuhara I, Malik KA, Miura M, Ohashi Y. 1999. Animal cell-death suppressors Bcl-xL and Ced-9 inhibit cell death in tobacco plants. Curr. Biol. 9:77578 Mittler R, Herr EH, Orvar BL, van Camp W, Willekens H, et al. 1999. Transgenic tobacco plants with reduced capability to detoxify reactive oxygen intermediates are hyperresponsive to pathogen infection. Proc. Natl. Acad. Sci. USA 96:1416514170 Mock HP, Heller W, Molina A, Neubohn B, Sandermnann H, Grimm B. 1999. Expression of uroporphyrinogen decarboxylase or coproporphyrinogen oxidase antisense RNA in tobacco induces pathogen defense responses conferring increased resistance to tobacco mosaic virus. J. Biol. Chem. 274:423138 Moon H, Lee B, Choi G, Shin D, Prasad T, et al. 2003. NDP kinase 2 interacts with two oxidative stress-activated MAPKs to regulate cellular redox state and enhances multiple stress tolerance in transgenic plants. Proc. Natl. Acad. Sci. USA 100:35863 Moseyko N, Zhu T, Chang HS, Wang X, Feldman LJ. 2002. Transcription proling of the early gravitropic response in Arabidopsis using high-density oligonucleotide probe microarrays. Plant Physiol. 130:72028

Annu. Rev. Plant Biol. 2004.55:373-399. Downloaded from arjournals.annualreviews.org by CAPES on 12/19/08. For personal use only.

ROS acts as a second messenger for the induction of defense genes in tomato plants in response to wounding, systemin, and methyl jasmonate. Plant Cell. 13:17991 Orozco-Cardenas ML, Ryan C. 1999. Hydrogen peroxide is generated systemically in plant leaves by wounding and systemin via the octadecanoid pathway. Proc. Natl. Acad. Sci. USA 96:655357 Ort RD, Baker NR. 2002. A photoprotective role for O2 as an alternative electronic sink in photosynthesis? Curr. Opin. Plant Biol. 5:19398 Pei Z-M, Murata Y, Benning G, Thomine S, Klusener B, et al. 2000. Calcium channels activated by hydrogen peroxide mediate abscisic signaling in guard cells. Nature 406:73134 Petit J-M, Briat J-F, Lobr aux S. 2001. e Structure and differential expression of the four members of the Arabidopsis thaliana ferritin gene family. Biochem. J. 359:57582 Polidoros AN, Mylona PV, Scandalios JG. 2001. Transgenic tobacco plants expressing the maize Cat2 gene have altered catalase levels that affect plant-pathogen interactions and resistance to oxidative stress. Transgenic Res. 10:55569 Prasad TK, Anderson MD, Martin BA, Stewart CR. 1994. Evidence for chilling-induced oxidative stress in maize seedlings and a regulatory role for hydrogen peroxide. Plant Cell. 6:6574 Purvis AC. 1997. Role of the alternative oxidase in limiting superoxide production by plant mitochondria. Physiol. Plant. 100:16570 Quinn J, Findlay VJ, Dawson K, Millar JB, Jones N, et al. 2002. Distinct regulatory proteins control the graded transcriptional response to increasing H(2)O(2) levels in ssion yeast Schizosaccharomyces pombe. Mol. Biol. Cell. 13:805 16 Rebeiz CA, Montazer-Zouhoor A, Mayasich JM, Tripathy BC, Wu S-M, Rebeiz C. 1988. Photodynamic herbicides. Recent

397

106.

97.

98.
Annu. Rev. Plant Biol. 2004.55:373-399. Downloaded from arjournals.annualreviews.org by CAPES on 12/19/08. For personal use only.

107.

99.

108.

100.

101.

109.

102.

110.

103.

111.

104.

112.

105.

113.

developments and molecular basis of selectivity. CRC Crit. Rev. Plant Sci. 6:385 436 Rizhsky L, Hallak-Herr E, Van Breusegem F, Rachmilevitch S, Barr JE, et al. 2002. Double antisense plants lacking ascorbate peroxidase and catalase are less sensitive to oxidative stress than single antisense plants lacking ascorbate peroxidase or catalase. Plant J. 32:32942 Roxas VP, Smith RK, Allen ER, Allen RD. 1997. Overexpression of glutathione S-transferase/glutathione peroxidase enhances the growth of transgenic tobacco seedlings during stress. Nat. Biotechnol. 15:98891 Sagi M, Fluhr R. 2001. Superoxide production by plant homologues of the gp91phox NADPH oxidase. Modulation of activity by calcium and by tobacco mosaic virus infection. Plant Physiol. 126:128190 Schopfer P, Plachy C, Frahry G. 2001. Release of reactive oxygen intermediates (superoxide radicals, hydrogen peroxide, and hydroxyl radicals) and peroxidase in germinating radish seeds controlled by light, gibberellin and abscisic acid. Plant Physiol. 125:1591602 Schopfer P. 2001. Hydroxyl radicalinduced cell-wall loosening in vitro and in vivo: implications for the control of elongation growth. Plant J. 28:67988 Singh KK. 2000. The Saccharomyces cerevisiae SLN1P-SSK1P two component system mediates response to oxidative stress and in an oxidant-specic fashion. Free Rad. Biol. Med. 29:104350 Solomon M, Belenghi B, Delledonne M, Menachem E, Levine A. 1999. The involvement of cysteine proteases and protease inhibitor genes in the regulation of programmed cell death in plants. Plant Cell 11:43143 Steinbeck MJ, Khan AU, Karnovsky MJ. 1992. Intracellular singlet oxygen generation by phagocytosing neutrophils

398

APEL

HIRT Carr R, Jhoti H. 2003. Oxidation state of the active-site cysteine in protein tyrosine phosphatase 1B. Nature 423:77377 Vanacker H, Carver TLW, Foyer CH. 2000. Early H2O2 accumulation in mesophyll cells leads to induction of glutathione during the hypersensitive response in the barley-powdery mildew interaction. Plant Physiol. 123:1289300 Vera-Estrella R, Blumwald E, Higgins VJ. 1992. Effect of specic elicitors of Cladosporium fulvum on tomato suspension cells. Evidence for the involvement of active oxygen species. Plant Physiol. 99:120815 Vernoux T, Sanchez-Fernandez R, May M. 2002. Glutathione biosynthesis in plants. In Oxidative Stress in Plants, eds. D Inze, MV Montagu. pp. 297311. London: Taylor and Francis Vianello A, Macri F. 1991. Generation of superoxide anion and hydrogen peroxide at the surface of plant cells. J. Bioenerg. Biomembr. 23:40923 Vranova E, Atichartpongkul S, Villarroel R, Van Montagu M, Inze D, Van Camp W. 2002. Comprehensive analysis of gene expression in Nicotiana tabacum leaves acclimated to oxidative stress. Proc. Natl. Acad. Sci. USA 99:1087075 Wagner AM. 1995. A role for active oxygen species as second messengers in the induction of alternative oxidase gene expression in Petunia hybrida cells. FEBS Lett. 368:33942 Whistler CA, Corbell NA, Sarniguet A, Ream W, Loper JE. 1998. The twocomponent regulators GacS and GacA inuence accumulation of the stationaryphase sigma factor sigmaS and the stress response in Pseudomonas uorescens Pf5. J. Bacteriol. 180:663541 Willekens H, Chamnongpol S, Davey M, Schraudner M, Langebartels C, et al. 1997. Catalase is a sink for H2O2 and is indispensable for stress defense in C3 plants. EMBO J. 16:480616 Wohlgemuth H, Mittelstrass K,

114.

115.

116.

117.

118.

119.

120.

121.

122.

in response to particles coated with a chemical trap. J. Biol. Chem. 267:1342533 Sweetlove LJ, Heazlewood JL, Herald V, Holtzapffel R, Day DA, et al. 2002. The impact of oxidative stress on Arabidopsis mitochondria. Plant J. 32:891904 Takeda T, Toda T, Kominami K, Kohnosu A, Yanagida M, Jones N. 1995. Schizosaccharomyces pombe atf1+ encodes a transcription factor required for sexual development and entry into stationary phase. EMBO J. 14:6193208 Tenhaken R, Levine A, Brisson LF, Dixon RA, Lamb C. 1995. Function of the oxidative burst in hypersensitive disease resistance. Proc. Natl. Acad. Sci. USA 92:415863 Tiwari BS, Belenghi B, Levine A. 2002. Oxidative stress increased respiration and generation of reactive oxygen species, resulting in ATP depletion, opening of mitochondrial permeability transition, and programmed cell death. Plant Physiol. 128:127181 Torres MA, Dangl JL, Jones JDG. 2002. Arabidopsis gp91phox homologues AtrbohD and AtrbohF are required for accumulation of reactive oxygen intermediates in the plant defense response. Proc. Natl. Acad. Sci. USA 99:51722 Torres MA, Onouchi H, Hamada S, Machida C, Hammond-Kossack KE, Jones JDG. 1998. Six Arabidopsis thaliana homologues of the human respiratory burst oxidase (gp91phox). Plant J. 14:36570 Tsugane K, Kobayashi K, Niwa Y, Ohba Y, Wada K, Kobayashi H. 1999. A recessive Arabidopsis mutant that grows enhanced active oxygen detoxication. Plant Cell. 11:1195206 van Breusegem F, Slooten L, Stassart JM, Moens T, Botterman J, et al. 1999. Overproduction of Arabidopsis thaliana FeSOD confers oxidative stress tolerance to transgenic maize. Plant Cell Physiol. 40:51523 van Montfort RL, Congreve M, Tisi D,

123.

124.

Annu. Rev. Plant Biol. 2004.55:373-399. Downloaded from arjournals.annualreviews.org by CAPES on 12/19/08. For personal use only.

125.

126.

127.

128.

129.

130.

131.

ROS Kschieschan S, Bender J, Weigel H-J, et al. 2002. Activation of an oxidative burst is a general feature of sensitive plants exposed to the air pollutant ozone. Plant Cell Environ. 25:71726 132. Zhang X, Zhang L, Dong F, Gao J, Galbraith DW, Song C-P. 2001. Hydrogen

399

peroxide is involved in abscisic acidinduced stomatal closure in Vicia faba. Plant Physiol. 126:143848 133. Zheng M, Aslund F, Storz G. 1998. Activation of the OxyR transcription factor by reversible disulde bond formation. Science 279:171821

Annu. Rev. Plant Biol. 2004.55:373-399. Downloaded from arjournals.annualreviews.org by CAPES on 12/19/08. For personal use only.

ROS

C-1

Annu. Rev. Plant Biol. 2004.55:373-399. Downloaded from arjournals.annualreviews.org by CAPES on 12/19/08. For personal use only.

Figure 4 Schematic depiction of cellular ROS sensing and signaling mechanisms. ROS sensors such as membrane-localized histidine kinases can sense extracellular and intracellular ROS. Intracellular ROS can also influence the ROS-induced mitogen-activated protein kinase (MAPK) signaling pathway through inhibition of MAPK phosphatases (PPases) or downstream transcription factors. Whereas MAP kinases regulate gene expression by altering transcription factor activity through phosphorylation of serine and threonine residues, ROS regulation occurs by oxidation of cysteine residues.

C-2

APEL

HIRT

Annu. Rev. Plant Biol. 2004.55:373-399. Downloaded from arjournals.annualreviews.org by CAPES on 12/19/08. For personal use only.

Figure 5 Different roles of ROS under conditions of (a) pathogen attack or (b) abiotic stress. Upon pathogen attack, receptor-induced signaling activates plasma membrane or apoplast-localized oxidases that produce superoxide radicals (O) that are 2 highly toxic and can help to kill the invading pathogen. On the other hand, O2 is rapidly dismutated into hydrogen peroxide, which, in contrast to superoxide, can readily cross the plasma membrane. Intracellular ROS levels increase due not only to extracellular production of ROS but also by downregulation of ROS scavenging mechanisms. Overall, ROS amounts increase to critical levels and induce programmed cell death (PCD). During abiotic stress, ROS production occurs mainly in chloroplasts and mitochondria at the sites of electron transport, increasing intracellular ROS amounts to toxic levels. The cellular response encompasses upregulation of ROS scavenging mechanisms to detoxify increased amounts of ROS.

Annual Review of Plant Biology Volume 55, 2004

CONTENTS
Annu. Rev. Plant Biol. 2004.55:373-399. Downloaded from arjournals.annualreviews.org by CAPES on 12/19/08. For personal use only.

AN UNFORESEEN VOYAGE TO THE WORLD OF PHYTOCHROMES,


Masaki Furuya 1

ALTERNATIVE NAD(P)H DEHYDROGENASES OF PLANT MITOCHONDRIA, Allan G. Rasmusson, Kathleen L. Soole,


and Thomas E. Elthon 23 41

DNA METHYLATION AND EPIGENETICS, Judith Bender PHOSPHOENOLPYRUVATE CARBOXYLASE: A NEW ERA OF STRUCTURAL BIOLOGY, Katsura Izui, Hiroyoshi Matsumura,
Tsuyoshi Furumoto, and Yasushi Kai

69 85

METABOLIC CHANNELING IN PLANTS, Brenda S.J. Winkel RHAMNOGALACTURONAN II: STRUCTURE AND FUNCTION OF A BORATE CROSS-LINKED CELL WALL PECTIC POLYSACCHARIDE,
Malcolm A. ONeill, Tadashi Ishii, Peter Albersheim, and Alan G. Darvill

109

NATURALLY OCCURRING GENETIC VARIATION IN ARABIDOPSIS THALIANA, Maarten Koornneef, Carlos Alonso-Blanco, and
Dick Vreugdenhil 141

SINGLE-CELL C4 PHOTOSYNTHESIS VERSUS THE DUAL-CELL (KRANZ) PARADIGM, Gerald E. Edwards, Vincent R. Franceschi,
and Elena V. Voznesenskaya 173 197 225 263 289 315

MOLECULAR MECHANISM OF GIBBERELLIN SIGNALING IN PLANTS,


Tai-ping Sun and Frank Gubler

PHYTOESTROGENS, Richard A. Dixon DECODING Ca2+ SIGNALS THROUGH PLANT PROTEIN KINASES,
Jeffrey F. Harper, Ghislain Breton, and Alice Harmon

PLASTID TRANSFORMATION IN HIGHER PLANTS, Pal Maliga SYMBIOSES OF GRASSES WITH SEEDBORNE FUNGAL ENDOPHYTES,
Christopher L. Schardl, Adrian Leuchtmann, Martin J. Spiering

TRANSPORT MECHANISMS FOR ORGANIC FORMS OF CARBON AND NITROGEN BETWEEN SOURCE AND SINK, Sylvie Lalonde,
Daniel Wipf, and Wolf B. Frommer 341

vii

viii

CONTENTS

REACTIVE OXYGEN SPECIES: METABOLISM, OXIDATIVE STRESS, AND SIGNAL TRANSDUCTION, Klaus Apel and Heribert Hirt THE GENERATION OF Ca2+ SIGNALS IN PLANTS,
Alistair M. Hetherington and Colin Brownlee

373 401 429

BIOSYNTHESIS AND ACCUMULATION OF STEROLS, Pierre Benveniste HOW DO CROP PLANTS TOLERATE ACID SOILS? MECHANISMS OF ALUMINUM TOLERANCE AND PHOSPHOROUS EFFICIENCY,
Leon V. Kochian, Owen A. Hoekenga, and Miguel A. Pi eros n
Annu. Rev. Plant Biol. 2004.55:373-399. Downloaded from arjournals.annualreviews.org by CAPES on 12/19/08. For personal use only.

459 495 521 537 555

VIGS VECTORS FOR GENE SLIENCING: MANY TARGETS, MANY TOOLS, Dominique Robertson GENETIC REGULATION OF TIME TO FLOWER IN ARABIDOPSIS THALIANA,
Yoshibumi Komeda

VISUALIZING CHROMOSOME STRUCTURE/ORGANIZATION,


Eric Lam, Naohiro Kato, and Koichi Watanabe

THE UBIQUITIN 26S PROTEASOME PROTEOLYTIC PATHWAY,


Jan Smalle and Richard D. Vierstra

RISING ATMOSPHERIC CARBON DIOXIDE: PLANTS FACE THE FUTURE,


Stephen P. Long, Elizabeth A. Ainsworth, Alistair Rogers, and Donald R. Ort 591

INDEXES
Subject Index Cumulative Index of Contributing Authors, Volumes 4555 Cumulative Index of Chapter Titles, Volumes 4555 629 661 666

ERRATA
An online log of corrections to Annual Review of Plant Biology chapters may be found at http://plant.annualreviews.org/

Das könnte Ihnen auch gefallen